New Bifunctional Perfluoroaryl Boranes. Synthesis ... - ACS Publications

Department of Chemistry, UniVersity of Waterloo. Waterloo, Ontario, Canada N2L 3G1. Department of Chemistry, South Road. UniVersity of Durham, Durham ...
2 downloads 0 Views 65KB Size
3244

J. Am. Chem. Soc. 1999, 121, 3244-3245

New Bifunctional Perfluoroaryl Boranes. Synthesis and Reactivity of the ortho-Phenylene-Bridged Diboranes 1,2-[B(C6F5)2]2C6X4 (X ) H, F)′

Thus, treatment of the known compound 1,2-(BCl2)2C6H411 with slightly more than 2 equiv of Zn(C6F5)2 led to smooth conversion to the desired product 1-H412 as shown in eq 1. While the first

V. Clifford Williams,† Warren E. Piers,*,† William Clegg,‡ Mark R. J. Elsegood,‡ Scott Collins,§ and Todd B. Marder| Department of Chemistry, UniVersity of Calgary 2500 UniVersity DriVe N. W. Calgary, Alberta, Canada T2N 1N4 Department of Chemistry, UniVersity of Newcastle Newcastle upon Tyne NE1 7RU, England Department of Chemistry, UniVersity of Waterloo Waterloo, Ontario, Canada N2L 3G1 Department of Chemistry, South Road UniVersity of Durham, Durham DH1 3LE, England ReceiVed January 11, 1999 Highly electrophilic boranes containing perfluorinated aryl groups are effective activators for olefin polymerization using d0 transition metal catalysts.1 To date, the most effective examples2 have been monofunctional boranes, but it has been suggested that bifunctional Lewis acids might offer advantages since the counterions formed from diboranes are potentially less coordinating than [RB(ArF)3]-.3 To this end, we have been developing routes to various diboranes of general formula (C6F5)2B-linkerB(C6F5)2 for testing as metallocene activators. Although such compounds are most relevant in the olefin polymerization arena, in a more general sense bifunctional boron-based Lewis acids are finding application in such diverse areas as organic synthesis,4 new materials,5 selective anion binding,6 and molecular recognition.7 Thus far, diboranes incorporating B(C6F5)2 units have been limited to those with one-carbon linkers,3a,8 which have some limitations and are, at any rate, best suited to binding hydride anions. Herein we report synthetic routes to diboranes containing bis-(pentafluorophenyl)boryl groups tethered by the two-carbon perprotio and perfluoro ortho-phenylene bridges. Recently, we reported the molecular structure of the base-free zinc compound Zn(C6F5)2,9 which we tested as a -C6F5 transfer agent.10 While it was not selective in reactions with BCl3, it is an effective reagent for converting -BCl2 units to -B(C6F5)2 groups. * Correspondence author. Telephone: 403-220-5746. Fax: 403-289-9488. E-mail: [email protected]. † University of Calgary. ‡ University of Newcastle. § University of Waterloo. | University of Durham. (1) Piers, W. E.; Chivers, T. Chem. Soc. ReV. 1997, 345. (2) (a) Li, L.; Marks, T. J. Organometallics 1998, 17, 3996. (b) Chen, Y.X.; Metz, M. V.; Li, L.; Stern, C. L.; Marks, T. J. J. Am. Chem. Soc. 1998, 120, 6287. (c) Yang, X.; Stern, C. L.; Marks, T. J. J. Am. Chem. Soc. 1991, 113, 3625. (d) Ewen, J. A.; Elder, M. J. Chem. Abstr. 1991, 115, 136998g. (3) (a) Jia, L.; Yang, X.; Stern, C.; Marks, T. J. Organometallics, 1994, 13, 3755. (b) Marks, T. J.; Jia, L.; Yang, X. U.S. Patent 5,447,895, 1995, Northwestern University. (4) (a) Reilly, M.; Oh, T. Tetrahedron Lett. 1994, 35, 7209. (b) Reilly, M.; Oh, T. Tetrahedron Lett. 1995, 36, 217. (c) Vaugeois, J.; Simard, M.; Wuest, J. D. Coord. Chem. ReV. 1995, 145, 55. (5) Fontani, M.; Peters, F.; Scherer, W.; Wachter, W.; Wagner, M.; Zanello, P. Eur. J. Inorg. Chem. 1998, 1453. (6) (a) Katz, H. E. J. Am. Chem. Soc. 1985, 107, 1420. (b) Katz, H. E. J. Org. Chem. 1985, 50, 2575. (c) Katz, H. E. Organometallics 1987, 6, 1134. (d) Herberich, G. E.; Englert, U.; Fischer, A.; Wiebelhaus, D. Organometallics 1998, 17, 4769. (7) (a) Nozaki, K.; Yoshida, M.; Takaya, H. Bull. Chem. Soc. Jpn. 1996, 69, 2043. (b) Nozaki, K.; Tsutsumi, T.; Takaya, H. J. Org. Chem. 1995, 60, 6668. (c) Nozaki, K.; Yoshida, M.; Takaya, H. Angew. Chem., Int. Ed. Engl. 1994, 33, 2452. (d) Katz, H. E. J. Org. Chem. 1989, 54, 2179. (8) Ko¨hler, K.; Piers, W. E.; Xin, S.; Feng, Y.; Bravakis, A. M.; Jarvis, A. P.; Collins, S.; Clegg, W.; Yap G. P. A.; Marder, T. B. Organometallics 1998,17, 3557. (9) Sartori, P.; Weidenbruch, M. Chem. Ber. 1967, 100, 3016. (10) Sun, Y.; Piers, W. E.; Parvez, M. Can. J. Chem. 1998, 76, 513.

two -C6F5 groups are incorporated relatively easily, the third and fourth substitutions require heating to 120 °C to ensure complete C6F5 transfer. 1H NMR spectra of the reaction in progress revealed the sequential nature of the substitution process eventually yielding 1-H4, characterized by a complex multiplet centered around 7.73 ppm for the aromatic protons of the symmetrical AA′BB′ phenylene backbone. In addition to full spectroscopic characterization, the molecular structure of 1-H4 has been determined (see Supporting Information). Boranes of general formula RB(C6F5)2 are prone to proteolytic loss of RH13 and are less Lewis acid than B(C6F5)3;14 furthermore, anions formed via methyl abstraction using such boranes (i.e., [R(CH3)B(C6F5)2]-) are unstable toward transfer of C6F5 groups back to the metal.8 Indeed, 1-H4 suffers from some of these problems in addition to being poorly soluble in aromatic solvents, which are usually employed in solution olefin polymerization processes. Accordingly, a more desirable ortho-phenylene-bridged diborane would include a fully fluorinated backbone which, like the parent borane B(C6F5)3, should be more stable to proteolysis15 and -C6F5 transfer and offer better solubility properties. The fully fluorinated analogue of 1, i.e., 1-F4, however, represents a more formidable synthetic challenge since the deactivating effect of four fluorine substituents precludes the use of a route analogous to that employed for preparing 1-H4. Not only were we unable to prepare 1,2-(Me3Si)2C6F4 but a model reaction between BCl3 and C6F5SiMe3 showed also that the silyl methyl groups preferentially underwent metathesis to boron instead of the required C6F5. We thus turned to the known mercury trimer [(C6F4)Hg]316 as a synthon for installing the 1,2-BX2 units on the tetrafluorobenzene ring. When this reagent is treated with an excess of BCl3, the desired 1,2-bis-(dichloroboryl)tetrafluorobenzene product is observed at early stages of the reaction, but is thermally unstable toward loss of BCl3 and production of octafluoro-9,10-dichloro9,10-diboraanthracene, which is the main product of this reaction.17 Eisch et al. have recently reported a similar condensation reaction involving unfluorinated diborane 1,2-(BCl2)2C6H4 at elevated temperatures.18 Evidently, this is a more facile process when the backbone is fluorinated.19 Fortunately, the reaction of [(C6F4)Hg]3 with BBr3 produces 1,2-bis-(dibromoboryl)tetrafluorobenzene, with only ∼5-10% of the diboraanthracene product (11) Kaufmann, D. Chem. Ber. 1987, 120, 901. (12) Experimental details and spectrocopic data for the new compounds prepared herein are given in the Supporting Information. (13) Parks, D. J.; Piers, W. E., Yap, G. P. A. Organometallics 1998, 17, 5492. (14) Deck, P. A.; Beswick, C. L.; Marks, T. J. J. Am. Chem. Soc. 1998, 120, 1772. (15) (a) Siedle, A. R.; Newmark, R. A., Lamanna, W. M.; Huffman, J. C. Organometallics 1993, 12, 1491. (b) Danopoulos, A. D.; Galsworthy, J. R.; Green, M. L. H.; Cafferkey, S.; Doerrer, L. H.; Hursthouse, M. B. Chem Commun. 1998, 2529. (16) Sartori, P.; Golloch, A. Chem. Ber. 1968, 101, 2004. (17) This compound and some of its derivatives have been fully characterized and will be the subject of a future publication. (18) Eisch, J. J.; Kotowicz, B. W. Eur. J. Inorg. Chem. 1998, 761. (19) For another example, see: Tschinkl, M. Schier, A.; Riede, J.; Gabbai, F. P. Inorg. Chem. 1998, 37, 5097.

10.1021/ja990082v CCC: $18.00 © 1999 American Chemical Society Published on Web 03/19/1999

Communications to the Editor

J. Am. Chem. Soc., Vol. 121, No. 13, 1999 3245 respectively) but, as is apparent from this view, the two borane trigonal planes are twisted out of the plane defined by the C6F4 backbone by about 39° to avoid unfavorable steric interactions. Furthermore, the two boron centers are tipped out of the C6F4 plane (B(1)-C(1)-C(6)-B(2) ) 19.9(2)°) to avoid each other. Despite the intraboron steric crowding, the B(1)‚‚‚B(2) distance of 3.138(2) Å is enough to accommodate small anions. Indeed, reaction of 1-F4 with potassium salts of F- and OH- in the presence of 18-crown-6 produces borate anions 3-F and 3-OH in which X- is chelated between the two boron centers (equation 3). Chelation of X- is implied by the two equivalent B(C6F5)2

Figure 1. ORTEP diagram of diborane 1-F4 (50% probability) with the F atoms removed for clarity. Selected bond distances (Å), nonbonded distance (Å), angles (deg) and dihedral angles (deg): B(1)-C(1), 1.568(2); B(1)-C(7), 1.565(2); B(1)-C(13), 1.568(2); B(1)‚‚‚B(2), 3.138(2); C(7)-B(1)-C(1), 121.72(13); C(7)-B(1)-C(13), 120.00(13); C(1)B(1)-C(13), 117.55(13); B(1)-C(1)-C(6)-B(2), 19.9(2); C(7)-B(1)C(1)-C(2), 41.8(2); C(13)-B(1)-C(1)-C(6), 39.8(2); C(19)-B(2)C(6)-C(1), 39.7(2); C(25)-B(2)-C(6)-C(5), 39.6(2).

in evidence (equation 2); this byproduct is easily removed in

workup. Conversion of 1,2-(BBr2)2C6F4 to 1-F4 by using Zn(C6F5)2 proceeds smoothly. It should be noted that attempts to prepare of 1-F4 directly from [(C6F4)Hg]3 and ClB(C6F5)213,20 were unsuccessful; thus, the use of the zinc reagent to institute the perfluorophenyl groups is essential. Diborane 1-F4 crystallizes as a toluene solvate;21 a view of its molecular structure (the F atoms and the toluene have been omitted for clarity) is presented in Figure 1 along with selected metrical parameters. Although somewhat distorted from ideal trigonal geometry, the two borane centers are essentially planar (ΣC-B-C ) 359.27(13) and 359.67(13) for B(1) and B(2), (20) Chambers, R. D.; Chivers, T. J. Chem. Soc. 1965, 3933. (21) Crystal data for 1-F4: C30B2F24‚C7D8, Mr ) 938.1; triclinic P1h, a ) 9.3690(6) Å, b ) 11.2168(7) Å, c ) 17.1652(11) Å, R ) 99.647(2)°, β ) 100.183(2)°, γ ) 104.082(2)°, V ) 1679.70(18) Å3, Z ) 2, T ) 160 K, Dc ) 1.855 g cm-3, λ ) 0.71073 Å, µ ) 0.20 mm-1. Data were collected on a Bruker AXS SMART CCD diffractometer. Refinement on F2 for all 7835 unique data (19422 measured, Rint ) 0.0255) gave Rw ) 0.0959 (on F2); conventional R ) 0.0360 on F values of 5931 reflections with Fo2 > 2σ(Fo2), difference map within (0.35 e Å-3. Programs were standard control software, SHELXTL (Bruker AXS Inc., Madison, WI, Version 5.1, 1998), and local programs.

units as indicated by the single (broad) peak observed in the non-temperature dependent 11B NMR spectra (at 14.0 and 2.8 ppm for 3-F and 3-OH, respectively) and the sharp, symmetrical 19F NMR spectra observed at all temperatures (213-313 K) for 3-F. Furthermore, the upfield shift of the para fluorine resonance relative to the meta signal for the C6F5 rings is indicative of a conversion from a neutral borane center to a partially anionic borate center.22 In 3-F, the µ-F ligand appears as a boronbroadened signal at -167.2 ppm (213 K) in the 19F NMR spectrum recorded in CH2Cl2. At room temperature, this resonance is nearly coincident with the signal due to the meta fluorines of the C6F5 rings (-165.0 ppm) but shifts upfield slowly as the temperature is lowered. The origin of the temperature dependence of the chemical shift for this resonance is not clear but, given the lack of coalescence behavior in the rest of the spectrum, we do not believe it is due to a fluxional process in which a nonbridging fluoride ion is exchanging rapidly between boron centers. Preliminary results suggest that other small anions such as Cland MeO- can also be accommodated by 1-F4. The incorporation of cations more appropriate for olefin polymerization applications is also feasible, and their use is currently being investigated. Acknowledgment. Financial support for this work was provided by the Natural Sciences and Engineering Council of Canada’s Strategic Projects program (STR192869) and the EPSRC (U.K.). Technical and in-kind support from Nova Chemicals of Calgary, Alberta, is also acknowledged. W.E.P. thanks the Alfred P. Sloan Foundation for a Research Fellowship (1996-2000). Supporting Information Available: Experimental and spectroscopic details for 1-H4, 1-F4, 2-F4, 3-F, and 3-OH, and tables of crystal data, atomic coordinates, bond lengths and angles, and anisotropic displacement parameters for 1-H4 and 1-F4 (PDF). This material is available free of charge via the Internet at http://pubs.acs.org.

JA990082V (22) See ref 8 and Horton, A. D.; de With, J. Organometallics 1997, 16, 5424.